Physica 12lA (1983) 399429 North-Holland
THERMODYNAMIC
Publishing Co.
PROPERTIES
OF INHOMOGENEOUS
FLUIDS
Michel BOITEUX Laboratoire
d’Ultrasons,
VniversitP Pierre et Marie
Curie,
75230 Paris Cedex
05, France
and John KERINS Department
of Chemical Engineering
and Materials Science, MN 55455, USA
University
of Minnesota,
Minneapolis,
Received 15 April 1983
A general method, the method of variation under extension, is presented for expressing the thermodynamic properties of an inhomogeneous fluid as functionals of the local number density, when given a density functional for the total thermodynamic grand potential of the fluid. The method is demonstrated in detail for the van der Waals square-gradient density functional and for the nonlocal density functional which arises in the theory of fluids with long-ranged pair potentials or in the mean-field theory of penetrable-sphere models. As specific examples, we consider the planar and spherical interface between two fluid phases, the line of contact of three fluid phases, the contact line between two surface phases and the planar interface between a solid and fluid.
1. Introduction
A simple example of an inhomogeneous equilibrium system is a sample of a single-component, subcritical fluid, in which a planar interface separates coexisting liquid and vapor phases, called phases b and LX,respectively. The thermodynamic properties of this inhomogeneous system depend on the properties of the coexisting bulk phases and on those of the planar interface’). For example, we must consider both the bulk-phase pressure p and the interfacial tension aOLfl when analyzing the mechanical work which can be done on the sample. On a microscopic level, this inhomogeneous fluid is most readily characterized by the density profile. With the coordinate x measuring distance perpendicular to the c$ interface (fig. la), the fluid density p(x) differs appreciably from its value in the c1or fl bulk phase only in the interface. A reasonable microscopic theory of the interface should provide a clear prescription for calculating interfacial thermodynamic properties, such as the tension @, given the appropriate microscopic information, say as contained in p(x). In fact in any microscopic theory of inhomogeneous fluid, one goal is to determine the thermodynamic properties 037%4371/83/000&0000/$03.00
0
1983 North-Holland.
400
M. BOITEUX
AND
J. KERINS
b
a
Fig. I. (n) A planar r/i interface: (b) a spherical r/j Interface; (c) a system of three bulk phases. 1. /j and T. which arc separated by two planar interfaces; (d) a system in which three bulk phases. 2. /j and ;‘, and three planar interfaces. r/l. z: and [P;. meet along a straight line of contact.
of the inhomogeneities sample of equilibrium
given their microscopic fluid as inhomogeneous
structure. We define a macroscopic if we can find two separated regions
in the sample, which contain isotropic fluid phases with a localized region of inhomogeneous fluid between the phases. For example, in contrast to the planar Z/I interface. we could imagine a bubble of s( phase separated from the surrounding /I phases by a spherical C$ interface (fig. lb); in this case the density p depends only on the radial distance r as measured from the center of the bubble. Or we could consider a fluid sample in which we find three regions, each filled with one of three distinct but coexisting bulk phases called 2, /I and 7. Now we might find the three phases separated by planar X/I and /?I) interfaces, with the fluid densities depending only on the coordinate s (fig. Ic). or, alternatively, we might find the three phases separated by three planar interfaces (shown as the planes xfi, /5; and r;~ in fig. Id), which meet along a straight line of contact’). In the latter case, where the fluid densities depend on coordinates .I-’ and .r2 perpendicular to the contact line, the line tension t is a thermodynamic quantity which should be determined in a microscopic theory, just as should the interfacial tensions a”‘. G’~‘,and ox,. In this paper we present a method for isolating and evaluating the thermodynamic properties of the inhomogeneities in an inhomogeneous fluid in the case that the equilibrium grand potential of the fluid is given by a density functional’). which is minimal with respect to variations in the local fluid densities. That is. we choose the density functional formalism as the framework for our microscopic theory of inhomogeneous fluid. The key to the method is an examination of the change in the grand potential Q of the fluid sample when the size of the sample
THERMODYNAMIC
is increased, practice
PROPERTIES
i.e., an examination determined
Accordingly,
while we do present
method
of the variation
the exact density functional
52 is usually
of variation
by example.
under
briefly
401
FLUIDS
extension.
for an inhomogeneous
by minimizing
under extension,
In an earlier
OF INHOMOGENEOUS
Of course
fluid is not known,
some approximate
density
some of the general
paper4) we were able to isolate
and
functional.
formulas
our main effort is to demonstrate
in
for the
the method
the line tension
of a
straight contact line by applying Noether’s theorem in the case of a van der Waals density functional. With the present method of variation under extension, we can not only reproduce this earlier result, but can also produce similar results for more general density functionals to which Noether’s theorem is not applicable. In section 2 we examine the variation under extension of the van der Waals density functional for a planar interface. In this simple case, we compute the variation in the grand potential, 652, by resealing directly the length coordinate. We introduce in section 3 a more elegant approach to determining 652 for a van der Waals functional through a functional differentiation with respect to the metric tensor, and then we apply the method in several different cases. This technique of functional differentiation is extended in section 4 to nonlocal density functionals, we discuss
and again the method is illustrated by several examples. In section 5 some of the limitations and further implications of the method of
variation under extension. results to classical results
2. The conventional
We also comment on the relationship in the theory of continuum mechanics.
of our present
analysis of variation under extension
Let us consider a single-component fluid at a fixed subcritical temperature T and at the associated coexistence chemical potential p(T), in which vapor or c1Phase is separated
from liquid or /I phase by a planar
interface.
We restrict our attention
to a macroscopic sample of the fluid contained in an imaginary rectangular cylinder, which has two faces of area A parallel to the interface and four edges of length 2L perpendicular to the interface (fig. 2). If the total grand potential of this sample is fi(A, L), then in a classical Gibbsian thermodynamic analysis, the grand potential per unit area or reduced grand potential Q(L) is L?(L)=7
S2”M L)
= -- 2pL + #‘p + 0(1/L),
with p the common value of the pressure in bulk phases a and ,!), and oUB the interfacial tension. Terms of order (l/L) are expected because of the fine size of the sample; even with L large, the fluid near the top or bottom of our imaginary sample container may be slightly different from the fluid in bulk a or p phase, respectively, because of the interface inside the container. Now imagine that we
M. BOITEUX
402
AND
-L-CL
Fig. 2. A sample area A.
extend
of the two-phase
our sample
J. KERINS
1
system in a rectangular
volume
by increasing
cylinder
the sample
from which we can conclude f+=lim
i
L-,X,
a(L)--lim?
C-10 t
2pL)c + 0(1/L)
2L and cross-sectional
height by an amount
top and bottom (0 < c 4 1) at fixed cross-sectional under this extension is 652, 652 = s2(L + EL) - Q(L) = (-
of height
area A. The variation
tL on
of Sz
)
(2)
that
1
(3)
Thus r~‘flis given by the asymptotic difference of the grand potential per unit area, Q(L), and the variation under extension divided by the scale of the variation, 60/t. If this two-phase system is described on a microscopic level by the well-known van der Waals or gradient theory of interface@‘), then the reduced grand potential
is taken
to be a functional
W)=Q[p;Ll=
j
of the equilibrium
density
dx{$(p)fc(p)(~~j.
profile p(x);
(4)
with c a positive function of p. The profile p(x) affords a minimum to the functional and necessarily satisfies the associated Euler-Lagrange equation.
(5) subject to the boundary the bulk GLor /I phases,
conditions that p(x) approach p’ or pB, the density in as x goes to + co, respectively. In the van der Waals
THERMODYNAMIC
PROPERTIES
OF INHOMOGENEOUS
FLUIDS
functional (4) the local density of grand potential is the sum of two the first tj(p(x)), which is the grand potential density in a uniform constrained to have density p(x) at the given T and ,u, and the c(p)(dp/dx)*, which is a positive contribution due to the inhomogeneity. goes to co, dp/dx goes to zero and $(p) goes to - p, the negative equilibrium pressure. For the given van der Waals functional (4), we could directly utilize the thermodynamic behavior (1) of Q(L) to write
403
pieces, system second As 1x1 of the known
@ = lim [Q(L) + 2pL] L-m
.
(6)
-02
In this integral expression for the tension aafi the integrand, although it does explicitly require the bulk-phase pressure p, is non-zero only in the interface. But now instead of this direct method, let us apply to the functional (4) the prescription for determining rr@embodied in the asymptotic difference formula (3) which follows from the method of variation under extension. We first extend the system by 2.5L in height, and then make a transformation of coordinates, (1 + c)y = x, to recover an integral with upper and lower limits of + L, respectively. L(l+G
s2(L + EL) =
dx
j-
-L(l
with b(x) = p(x + a). #iqx)=p(x)+
(
$
{W+cCjo)‘)
+c)
68 is now calculated by expanding p’(x) about p(x),
>
cx+O(c2),
(8)
with the result that
(9)
404
M. BOITEUX
Given
the asymptotic
The integral
difference
expression
the equilibrium
profile
AND J. KERINS
formula
(3) for a’” we then have
for a”’ can be further in the form
in the integrand of (IO). and formula for the tension,
then
simplified
integrating
by parts
by first using eq. (5) for
to find a well-known
(12)
While we have again expressed only in the interface,
the present
cr’fi as the integral integral
of a function
which is non-zero
(12) differs in several important
respects
from the earlier integral formula (6). The integrand in (12) is simpler than that of (6) and could be more convenient for numerical evaluation of crXB.Indeed, the density latter integral expression (12) for 0 ‘1 is valid only if p(s) is the equilibrium profile; the difference in aXB as calculated from (6) and (12) could be used as an estimate of the numerical error. More importantly, it is obvious from (12) that rrrB > 0, in accord with thermodynamic stability, but no such conclusion is readily apparent
from (6) since the structure
of G(p) is not specified.
The method of variation under extension may be applied to quantities other than the grand potential per unit area. For example, if N is the total number of particles
in our sample,
then choosing
the Gibbs
dividing
surface at x = 0 we may
write
where
r is the adsorption
N =
at the interface.
On a microscopic
J
d.v p(s)
and by following
the steps of (7))( 12) we deduce
level we have
(14)
that
THERMODYNAMIC
PROPERTIES
OF INHOMOGENEOUS
FLUIDS
405
From a thermodynamic point of view (7@ is the quantity of fundamental interest. Because we started with the grand potential of the system, o@(T, p) is the thermodynamic potential for the interfacial phase, i.e., dazP=s”fidT+r
(16)
dP,
where s’fi is the interfacial entropy density. Furthermore, a@ is independent of the choice of Gibbs dividing surface whereas r and s@ are not. In terms of the microscopic description, if p,(x) is a solution of the profile eq. (5), then so is p*(x) = p,(x + x,,), x,, a constant; but whereas the surface tension (12) of the two solutions will be the same, the adsorptions will be different r, # Tz for the same choice x = 0 of the dividing surface. Moreover, only in the integral expression (12) for alp did we explicitly use to advantage the fact that p(x) is an equilibrium density profile. On a macroscopic or thermodynamic level, the fundamental equation for the method of variation under extension as applied to the planar interface is (3). The method is useful, however, only because for the given van der Waals functional (4) the variation 6Q can be explicitly calculated. The key step in this calculation of 652 was resealing the length coordinate (7) which effectively transforms a variation at the boundaries of the system into a variation over the interior of the system. But now the variation over the interior depends on the structure of the system, and in particular on the structure of the interfacial inhomogeneity. It is just because of this dependence of 652 on the structure of the inhomogeneity that we can isolate the thermodynamic properties of the interface through the macroscopic equation (3). In the next section we show how a variation on the boundary of an inhomogeneous fluid sample, for which the reduced grand potential is given by an appropriate van der Waals density functional, may generally be transformed into a variation over the sample interior, and hence how the method of variation under extension may be used to determine the thermodynamic properties of the inhomogeneities as functionals of the density.
3. The variation under extension
of a general van der Waals density functional
Suppose we consider a macroscopic sample of a C-component fluid in which there is a d-dimensional inhomogeneity. That is, the local densities p’, i= 1,2 3 . . 3 C, are functions of d coordinate variables xX, s = 1,2, . . , d. The reduced grand potential of this sample, Q, has the form Q(L) = o,Ld+
o,_,Ld-’
+. . . + 0,L + o,+
G(l/L)
.
(17)
L is a macroscopic length associated with the d-dimensional region S(L) over which Q(L) is defined; (0, is the density of grand potential in the i-dimensional
M. BOITEUX
406
phase
as defined
homogeneity
in a Gibbsian
is ri-dimensional,
AND
thermodynamic
example,
(n-
of the (T-dimensional
d). In the case of the planar
where d = 1 and J=
per unit interfacial
analysis’).
Although
the fluid itself may be &dimensional,
d 3 d; Q(L) is the total grand potential of dimension
J. KERINS
3, 2L is sample
liquid-vapor height; B(L)
fluid per unit “area” interface
taken
P-phase,
bubble
in which the a-phase
as the radius is centered;
(fig. 2), for
is the grand potential
area; w, = - 2p and (II,, = 0’8. For the bubble
d = a = 3. L is conveniently
the inas long as
(fig. 3a) where
of the spherical
volume
of
Q(L) is n”(L), the total grand
potential; oj = 471/3 p”, where p” is the pressure of the bulk /I phase; o> = tc), = 0 and cc)”is the excess grand potential associated with this spherical inhomogeneity. In the case of a straight L is taken
three-phase
as the radius
(fig. 3b), where d = 2 and a = 3,
line of contact
of the cylinder
enclosing
the contact-line
sample;
Q(L)
is the grand potential per unit length of the contact line; (o? = - 7cp. where p is the common value of the pressure in each of the bulk phases (r, fl and y); tc), = @ + afi + 0%;’ is th e sum of the surface tensions and (ti,) = 7 is the line Our aim is to isolate the contribution to Q(L) from the individual o,L’. the variation of O(L) under extension. i =o, 1,. , rf by examining In a van der Waals theory, the reduced grand potential Q(L) is an integral over tension.
S(L)
of a function
dimensional) Q(L)
gradients = L?[p’; L] =
The p’(x) are Euler-Lagrange functional
Y depending VP’. i = 1,2,
on
the densities
. , C; e.g. in Cartesian
1
p’(x)
and
their
(18)
3
and necessarily are solutions of the densities, SQ/6p’(x) = 0, associated with the van der Waals
the extension
652= Q(L + CL) -L?(L)
L + L + CL, from (17) one finds that
= t 1 iw,L'+
C(l/L))
1-I
a
(cl-
coordinates
dx ‘f’(p’(x), BP’(X))
equilibrium equations,
(18). Under
only
(19)
b
Fig. 3. (a) A bubble of r phase inside a b-phase sphere of radius L. (b) A contact line sample enclosed in a circular cylinder of radius L,in which the axis of the cylinder is parallel to the straight three-phase line of contact.
THERMODYNAMIC
so that the contribution
A d-l
=
PROPERTIES
OF INHOMOGENEOUS
of odLd to Q(L)
Q(L)-ljlg
=y
!=I (
may be removed
variation
of A,-,
I
for the present
under extensions
to get A,-,,
0;L’.
l-6
(Here and below we neglect terms of order are of no consequence
407
FLUIDS
l/L in writing
discussion.)
equations;
We could
and remove the contribution
such terms
now consider of od_, Ld-
the
’ from
A N_, to get A,_,, and so on to A, = co,; given the set of d + 1 quantities Sz, , A, we could clearly extract the quantities w,L’, i = 0, 1, . . , d. In A (,- ,, applying potential,
this procedure to the van der Waals functional (18) for the grand we also want to explicitly make use of the fact that the functional is
minimal with respect to variations in the p’(x). Let us write the functional Q[ p’; L], given previously (18) in Cartesian coordinates, in a general curvilinear coordinate system x, where x = (x’, x2, , x”> is the contravariant coordinate vectors),
Q[ pl; L] =
dx ,/m s S(L)
y ( P’>~,P’>cc,,) 3 (21)
4P’h)
W(x) = F.
The gradient of pi, Vp’ = (a,~‘, &p’, . . . , adpi), is a covariant vector. g,,(x) is the metric or fundamental tensor for the given coordinate system (see refs. 8-10 for its properties; any inner product of gradients, such as Vp’ - VP’, depends on the metric tensor). m is the Jacobian of the Cartesian to curvilinear coordinate transformation, and is simply related to the determinant of g,7,, as g(x) = ldet g,s,(x)l. In the general curvilinear coordinate system, we should properly regard the reduced grand potential as a functional of the densities and the metric tensor, i.e. s2 = !2[ p’, g,,; L]. Let us now extend the system by increasing L to L + CL, and then make a change of coordinates, x = x(y, 6) such that S(L + CL), the domain of integration after extension in the x variables, is mapped onto a domain of integration in the y variables that is S(L). For example, in Cartesian coordinates x is related to y by a simple dilation or resealing transformation; indeed the transformation x = x(y, t) represents the simple dilation in terms of the general curvilinear coordinate system. After such a coordinate transformation or resealing, we can evaluate the variation 6Q by a functional Taylor expansion, since the domain of integration for both Q(L) and sZ(L + CL) is S(L). Obviously we must account for the
M. BOITEUX
408
variations
induced
by resealing
632
f2
AND J. KERINS
in both the densities
632 &!?\,(X)&J‘(Y) + . W(x)b’(JJ)
&L(X)W(Y)
and the metric.
-i &
Sp’(x)dp’(y)
.
I
(22) where for the given change
of variables
d.u’(y, c) ?.u’(y. t )
dg:,,(y) =7
~
R,,
8,,’
(x (Y, ( 1)- &,.(Y 1
and
(23) = P’(X(Y? ( 1) - P’(Y)
MY)
We have adopted the summation convention of tensor analysis’), in which any repeated index is to be summed over; this applies to the index of the (contravariant) coordinate vector .Y(s’; s = I, 2, , d) and to the indices of the metric tensor (K,,; s. t = I,?, , d). It also applies to the index i on the densities o’(x), although in this case i is simply an index for the set of the C scalar densities and does not imply any vector properties. Only because the densities are scalar quantities does it make sense to refer to the gradient V/I’. The coordinate transformation
can be expanded
x(y. <) =y
in terms of (
+ Uj(y) + (“(C2).
since for ( = 0, x = y, the identity can then be evaluated k:,,(Y)
= 6 k,,~,v”
6p’(y)
= ~[y1’cQl’]$_
Since these variations
+ R,$,V” C’(f’)
(24) transformation.
+ ‘1”~,gJ = c[q *Vp’]
The variations
+ c’(e)
fig,, and bp’ (23)
.
(25)
+ C’(c?).
6x,,, and 6p’ are of order
(26)
c. we may rewrite
6~ (22) as
(27) But recalling
that Q is minimal
with respect to the variations
in the densities
(i.e.
THERMODYNAMIC 652/6p’(x)
= 0), we
PROPERTIES
OF INHOMOGENEOUS
FLUIDS
409
have
(28)
where T”’ = Ts’(pi, aspi, g,,) is defined in the following equations,
(29) The (contravariant) tensor g” is the inverse of g,,; g”‘g,,.= S:, where S:, is the Kronecker delta. T” is commonly called the momentum or energy-momentum tensor in field theories’.“). Knowing the variations Sg,, (25) and 652 (28) we could calculate A,_, through (20), then consider the variation of Ad_, under extension to get A,_,, and so on. Of course in evaluating 6A,_, it will not be true that 6Ad_ J&pi = 0 (cf. eq. (27) for AC!), but this does not cause difficulty; the method of variation under extension is not limited to functionals which are extremal. Usually in field theories, as Landau and Lifshitz remark”), “. . the metric tensor has no independent significance and the transition to curvilinear coordinates occurs formally as an intermediate step in the calculation of T”“. In the method of variation under extension, however, g,, has assumed a real role in the calculation of 652. We have implicitly assumed that the fluid with a d-dimensional inhomogeneity fills the entire J-dimensional space, although we isolate a finite sample in an imaginary container when defining Q(L) and Q(L + CL). Thus Q[p’, g,Y,;L] and Q[p’,g,,; L + EL] are functionals of the same profiles p’ and metric g,,,, but the former functional is defined on the domain S(L), while the latter is defined on S(L + CL). This is just as in the discussion of the planar interface in section I; the Euler-Lagrange equation (5) for the profile p(x) was independent of L, as were the boundary conditions. We used a coordinate transformation in examining the variation of the grand potential under extension so that s1(L. + CL), like Q(L) itself, was defined on the domain S(L). Q(L + CL) was transformed into a functional over S(L) of the new functions p’ + 6~’ and g,7,+ 6g,,. The variations 6~’ and & in essence map the boundary extension of S(L) to S(L + CL) into a variation over the interior of S(L), and as such have real significance. Or, in less precise but perhaps more familiar terms, when we “compress” our sample from S( L + EL) to S(L), we induce not only a direct local change 6p’(x) in the (density) fields, but we also induce a local “strain” 6g,, in the volume, which is coupled to a local “stress” T”‘. Since the grand potential is extremal with respect to variations in the densities, sZ(L + CL) is insensitive to direct local changes 6~’ and only the
410
M. BOITEUX
“strain-stress”
variations
AND J. KERINS
are important
in 6.0. We discuss
of view further
in section
5.
At this point
the main
steps in the method
clear, and rather we now consider suppose gradient
than pursuing three
a general
specific
that the function form
of variation
prescription
examples.
this mechanical under
extension
functional
are
. d - 1,
for LI,, i = 0, 1,
For each of these examples
Y in the density
point
we will
(20) has a van der Waals
v’(P’. (7,P’.g,,,) = Ii/(P) + ‘.,~(/‘)‘?,p’g,,i’,p~
= Icl(P)+ “,r(Pm’w~
(30)
$(p) and c,JP) are functions of the C number densities in this C-component system. Although the summation convention does apply to the indicesj and k of c(.j,k=
1,2, . ..) C), these indices
do not imply any tensor
character
for the c,~.
i) The struight line of‘ contuct of fhrw .&id phmes’,“) Given an inhomogeneous fluid in which three bulk phases (r, b, y) and three planar two-phase interfaces all meet along a straight line of contact (fig. Id), let us isolate a macroscopic sample of the contact line in a circular cylinder (fig. 3b) of radius L with the axis of the cylinder parallel to the line of contact. Suppose that the Gibbs dividing surfaces, shown as the planes a/?, ~7, and 811 in fig. Id. have a common line of intersection, the contact line, along the axis of the cylinder. The grand potential per unit length of the contact line has the form (with d = 3 and d = 2). (31) where p is the common value of the pressure in each of the three bulk phases. 0 h’ is the interfacial tension of the ~2 interface (K,? = zj?,pp, cc;))and 7 is the line tension.
On the basis of eqs. (20) and (3l),
(32)
A,, = z In the van der Waals
c?[p’; L] =
density
dx J&k s .VL)
functional
for Q(L)
given earlier
in eq. (21)
P’. (‘,P’, g,,) .
with the function ‘P as in (30), the equilibrium profiles p’ depend on the two-component vector x, for which the domain of integration S(L) is a circular
THERMODYNAMIC
PROPERTIES
OF INHOMOGENEOUS
FLUIDS
411
(I
cii3 X2
‘\
/’
/’
X'
mlny
P
I
V
L
Fig. 4. The domain of integration S(L) for x’ and x2 is a circular cross-section of radius L of the contact-line sample. The point (m, n) denotes an alternative location for the contact line associated with an alternative placement of the Gibbs dividing surfaces.
cross section of the cylindrical sample, perpendicular to the contact line and with area XL* (fig. 4). For the present calculation of A, and do, we choose a Cartesian coordinate system (xl, x2) in which the coordinate origin is at the center of S(L). Upon variation of the length L, the appropriate resealing coordinate transformation (24) is simply x(y, 6) = (1 + 6)~ (i.e. q = y) which implies that sg,, = 6gg22 = 2lz) 6gg12 = agg2r= 0. Furthermore,
for !P (30) in Cartesian coordinates,
T1’=$(P)+
(33) it follows from (29) that
cj~[a2pja2pk--alpjalp~,
T” = T2’ = - 2Cjka,Pja2Pk,
(34)
T22= $(P) + cjk[d#jalpk - a2pja2PT. From eq. (28) for a&?,we can then conclude that
1imE= c+o 2c
dx +(p). s S(L)
(35)
This in turn implies that A, =
dx Jgx)Dd~‘,
aspivgrJ
s
(36)
s(L) with Dr = cjk&p’g”‘a,pk = c&”
Vpk .
(37)
With the line tension and one-half the surface tension isolated in A, (32), we now consider the variation of A, under extension, (38)
412
M. BOITEUX
The functional
derivative
AND J. KERINS
of d, with respect
to g,, can be evaluated
Q with d, and Y with D, in (29); for the particular problem,
one can show that (6d,/6g,,)&,
6p’(y) (26) is straightforward
equals
0. Calculation
(recall q = y), as is the calculation
follows from the macroscopic
equation
by replacing
form of D, (37) of the present of the variation of cSd,/Gp,. It then
(32) for d, and (38) for the variation
cid,
that
zz
s
dx (!,[s ‘c,Jp
’ - vp “1 + ?z[_~2c,,~p1. vp “1
(39)
S(L)
The integrand
on the r.h.s. of (39) is clearly a divergence,
and we could transform
C a”‘(L/2) to a contour integral around the boundary of S(L). In such a contour integral with L very large, we can indeed recognize factors associated with the individual tensions ah-’ (cf. the earlier surface tension formula of eq. 12). Finally we can produce, tension r,
on the basis of (32) an integral
expression
s
=
dx s ‘?,[c,kVp ’ * V/j “1
for the line
(40)
s, I , Alternatively.
we could
isolate
the surface
tension
contribution
to iiQ/c in (35)
(41)
and
then
deduce
from
(32) and
(36) the more
familiar
formula
(in Cartesian
coordinates)
=s dx
~,J’P’.~P’-
iti
+p])
(42)
.S(71
The two expressions for T, (40) and (42) are equivalent as can be shown by using the fact that the equilibrium densities p’(x) satisfy the Euler-Lagrange equations associated with the density functional Q[p’; L] (i.e. the equations ?Q/6p’ = 0). Eq.
THERMODYNAMIC
PROPERTIES
OF INHOMOGENEOUS
FLUIDS
413
(42) is particularly useful in the numerical computation of the line tension r *,“). Our assumption that the Gibbs contact line coincides with the axis of the cylinder enclosing the three-phase sample is arbitrary. Suppose that we had a different placement of the Gibbs dividing planes and line, one where the contact line intersects the cross section S(L) at the point (m, n) (fig. 4). With this choice we must write a more general surface tension sum c oK’hKiin (31); h””is the length of the ~1 interface in S(L). Although the lengths h”’ are functions of L, m and n, under the extension of L to L + CL, the variations 6h”” are (43) Given this behavior (43) for the variations, there is essentially no difference in the calculation of d, and d,. But then in eq. (39) or (41) we have an (asymptotic) integral expression for the surface tension sum, which is clearly independent of the choice (m, n). This implies that 0 = lim 1 oK’:L- 1 o”h”(L, m, L-a
n)],
(44)
[
independent of m and n. Eq. (44) is equivalent to the well-known Neumann triangle conditions for thermodynamic mechanical equilibrium at the contact line’2?. ii) The contact line between two surface phases4,13) Let us consider the meeting at a straight contact line of two distinct structures for the planar interface between bulk phases c( and b (fig. 5). For example, we could imagine an interface at which one of the species is strongly adsorbed in contact with an interface at which the same species is weakly or even negatively adsorbed. A macroscopic sample is isolated in an imaginary rectangular cylinder of width L, height L’ and length L”, whose axis is parallel to the line of contact. In a Gibbsian thermodynamic analysis of this system, there is one dividing plane (shown at height x; in fig. 5~); this plane is separated into two sections by a straight line of contact. The section to the left of the line has area L”w, 0 < w < L and is associated with the first (I) interface structure of tension a;fi; the right-hand section has area L”(L - w) and is associated with the second (II) interface structure of tension rr$“. If the common value of the bulk-phase pressures is p, then the excess grand potential per unit length is a(L) Q(L) =
QL, L’, L”) + pLL’L” = 0;‘L + [(0$ L”
O;')W + 51,
(45)
where T is the line tension of the interfacial contact line. Although the inhomogeneity in this case is two-dimensional, d = 2, by removing the bulk-phase pressure beforehand in 0 we have effectively fixed o2 = 0. The density functional
M. BOITEUX
414
AND
J. KERINS
b
Fig. 5. (a) The straight line of contact between structures I and II for the planar I[{ interface; (b) A sample of the two-phase system enclosed in a rectangular cylinder, whose axis is parallel to the straight line of contact: (c) the interfacial Gibbs dividing plane at height _r’,.
for a is still that of (21) except that in the function Y (30) which appears in the integrand of Sz[ p ‘, L] we should replace tj (p) by J(p) = $ ( p) + p. If we choose a Cartesian
coordinate
the x’-axis
parallel
perpendicular
system
to the contact
.?-direction (with the _r2 = x’/(l + L)) we have 6g22 = 2t , Applying
interface
is perpendicular
in a cross-section
line (fig. 5c), then in the variation
subsequent
resealing
behavior
A, = a;“L =
equations
and
S(L, L’) which
transformation
of Q(L)
is
in the
J’ = .Y’ and
Sg,, = 6g,: = sgz, = 0.
the appropriate
thermodynamic
(.u’, _v~) in which the xi-axis
to the planar
(46) of variation
(45) we conclude
d.v r&p) + ‘;J&P’?,P~
(28) and
(29) for the given
that
- &‘??pk]
(47)
and
A0 = [(a;/ - (T;‘$v + T] =
s
dx 2c,,&p j&p k .
(48)
S(L.L’)
Now we observe O(L)
= o$L
that we could have written +
[(@ - &)W + t]
)
the thermodynamic
formula
(45) as (45’)
THERMODYNAMIC
PROPERTIES
OF
INHOMOGENEOUS
FLUIDS
415
where M”= L - w has the same degree of arbitrariness as w. We would then find the same expression (47) for d,, but now equal to [T$L. This leads to the conclusion that opfl= o$, a necessary condition of thermodynamic equilibrium for the interface, and hence to the conclusion that d, = z. iii) Spherical bubble As the last example in this section, we wish to consider a bubble of cc-phase located at the center of a spherical sample of b-phase, in which the sample radius is L (fig. 6). In this case L?= d = 3 and !2(L) is the total grand potential a(L),
where p” and pB are the pressures in phases a and p, R is the radius of the Gibbs dividing surface, and r~x0 is the interfacial tension. In order to make connection with some earlier work’4.‘5) on th e van der Waals theory of spherical interface, we choose a spherical coordinate system (fig. 6) x’ =r,
x2= 4,
x3=0
(50)
and restrict ourselves to a one-component functional (21) may be written
system. Then the van der Waals
(51) In examining transformation
the variation of G? under extension in L (with the resealing affecting only the radial variable r) we have
&L, = 2&V, &,, = 0,
Fig. 6. The spherical bubble is surrounded
s z t, s, t = 1, 2, 3 ,
coordinate system by fl phase.
for the spherical
(52)
sample
of radius
L in which
an a-phase
416
M. BOITEUX
which leads from (28)
AND J. KERINS
(29) and (49) to the conclusion
that
(53)
We can further
deduce
that
(p!‘-p’)$R’+rr”4nR’ 1 Thus
we have a simple
integral
expression
for the contribution
to 52 from
the
bubble-inhomogeneity, both the interior a phase and the spherical c$ interface. At the given T and p for this one-component system, the phase /I exists as a bulk (stable or metastable) equilibrium phase and so has a well-defined pressure p”( T, p). In a Gibbsian thermodynamic analysis, p’ is the pressure of a homogeneous reference phase c( with the same T and p as the bulk p phase, but in general there is no clear thermodynamic prescription for explicitly evaluating p’; in the present van der Waals theory, however. both p’ and p” are determined by the function Ic/(p). Accordingly, the choice of the reference pressures p” and p ii is independent of the radius R of the Gibbs dividing sphere. If we differentiate (54) with respect to R we find the thermodynamic Laplace equation’.‘“) (55) since the r.h.s. of (54) is independen write
‘55) in (54) we may
(56)
Only the particular combination of p’, p”, cXp and R given on the 1.h.s. of (54) equals the total excess grand potential associated with the bubble. and as such is independent of R. The arbitrariness as to how this total excess is assigned to the interior reference phase a or the spherical c$ interface leads to a formal dependence of the surface tension (56) on the choice of R. If we choose R,, as the radius where [daXB/dR] = 0, then (55) reduces to the classical Laplace equation of capillarity and R,, is called the radius of the surface of tension. In this case a’0
THERMODYNAMIC
and R, are unambiguously
PROPERTIES
determined
OF INHOMOGENEOUS
417
FLUIDS
by (55) and (56)
and 4 I$ dr r2c( P)(dp /dr)2 %=
(P-P9
1 “3
(57b)
Thus we have simple integral expressions for the radius of the surface of tension and the surface tension at that radius, which depend only on the reference states LXand /3 and on the equilibrium density profile p(r). These three examples demonstrate how the technique of variation under extension can be applied in density functional theories of the van der Waals type to deduce nontrivial functional expressions for the thermdoynamic properties of the inhomogeneities, e.g. (42), (48) or (54). Moreover, certain conditions of thermodynamic equilibrium, e.g. (44) or (55), follow directly from the invariance of functional expressions to the choice of Gibbs dividing surfaces. In the next section we extend the method of variation under extension to a particular class of nonlocal density functionals.
4. Nonlocal
density functional
We now examine the variation under extension of the reduced grand potential 52 of an inhomogeneous fluid sample in the case that C! is given as a nonlocal functional of the equilibrium densities p’(x), Q[P’; Ll
= dx dv &it%?i%,d~~~)~‘(-W(y) s
s
+
s S(L)
dx Jg(x)W.
(58) Here h,,(x, y) are functions
of x and y, and Ic/(p) is a function
of the densities
p’.
Such nonlocal density functionals appear, for example, in the mean-field theory of penetrable-sphere models3.12), and in the theory of fluids with long-ranged pair potentialsI 19). ‘/I(P) is that part of the local density of grand potential which depends only on the densities pi at X. Depending on the model it might be the grand potential of an ideal gas3.12), or of a reference hard-sphere systemI 19). The interaction of fluid particles at x with those at y is responsible for the nonlocal interaction kernels hj(x, y). The domain of integration for x and y in the first term on the r.h.s. of (58), the nonlocal piece, must be specified. The point x may be anywhere in the region S(L)
418
M. BOITEUX
which is determined
by the shape of the macroscopic
simply one portion. which occupies
AND J. KERINS
isolated
the entire
in an imaginary cj-dimensional
with fluid particles
assume,
that the values of the interaction
if x and y are greater neighborhood
than
of radius
at points
a distance
< of x ES(L)
Since the sample is
of an inhomogeneous
space, fluid particles
may still interact however,
sample.
container,
at points
s. which lie outside
5 apart.
kernals Thus
fluid x ES(L)
of S(L).
We
/z,~(x,J) are negligible only
need be considered,
points
y within
and the domain
a of
integration for y in (58) may be restricted to the region S(c) which contains S(L) but is larger in extent by a factor of 4, i.e. t = L + [. When we examine the variation of R(L) under extension, we implicitly will be varying both S(L) and S(E). Note also that the equilibrium profiles p’(x) necessarily satisfy the Euler-Lagrange equations associated with the functional (58) in the limit of infinite sample size. That is, for any x in S(-,), where S(x) is the limiting region for S(L) in the limit that L goes to CC. 0~
a*
lx2
]im -==+ &l’(x)
IT.L-rL
&7’(x)
s
dy h(x>y)
+ ~,(~.Y)IP~(Y),
subject to the appropriate boundary conditions at infinity. i.e.. on the boundary of S(a). For this reason the variation of the functional Q(L) of (58) with respect to the densities p’(x). XES(L) (60)
is not quite 0 because of the finite or truncated domain S(L). Nevertheless, in all cases below we will set SQ/6p’ to 0 since the truncation error, which originates in the difference between S(L) and S(x), vanishes in the limit that L and E go to CD. In a general actually kernals
curvilinear
a functional
coordinate
of the densities
system, the nonlocal functional (58) is the metric tensor g,, and the interaction p’,
hiA;
Q(L) = Q[P’. g,,, h,r;
Ll
(61)
When examining the variation of C?(L) under extension in L, we must account for the variations 6~’ in the densities, fig,, in the metric and 6h+ in the interaction kernals. Thus (cf. eq. 22), 6Q
. -----p'(x)+ W(x)
"nk,Wj+j &v(x)
.%I., x
SC/~) +
higher-order
dx
terms in the variations.
j
dy
{~6h,(v)j
XL)
(62)
THERMODYNAMIC
PROPERTIES
OF INHOMOGENEOUS
FLUIDS
419
As always, the grand potential is insensitive to local variations in the density since &2/6p’= 0, so the first term on the r.h.s. of (62) vanishes. The earlier formulas still apply for evaluating the functional derivative with respect to the metric (29) and for the variation in the metric (25). The variation 6hjk can be calculated for the earlier resealing coordinate transformation (24)
(63) and the functional derivative &2/6hjk is also available g
=
&b”(Y).
(64)
/k
We may now write the final result, based on (62), in the form limE= c+o t
dx dY ,/&%(%‘(x)~~(v) s s S(L) X S(L) -Mx) + ~MY) axs ays
+
+
[11 (x ) ’ V.&j/c + ‘I (Y) ’ V&d
s dx
1
J~(X)J&)PW(X) +gT%gtsl~
(65)
S(L)
With this result we could calculate A,_,, then determine A,_, and so on. Again though, rather than proceeding with a general formalism, we turn instead to four specific examples. i) The planar liquid-vapor interface in a van der Waals fluid Let us imagine a one-component two-phase system with a planar interface (fig. 2), in which the individual fluid particles interact pairwise with a potential 4(r) of the type proposed by Kac”) (for C?= 3)
M. BOITEUX
420
If we adopt
the properly
scaled coordinates
limit of small y, the grand (approximate)
functional
potential
x = YYand take the (van der Waals)
per unit area Q(L)
for this system
has the
form
L
YQ[P, Ll=
AND J. KERINS
/.
s
L
d.r e -1’ %(.‘oP(.r) -L
I
(67)
IT
with the Cartesian coordinate .Yas in fig. (2). Note that the length L has also been scaled by the inverse length y, and we have implicitly assumed E > L. The dependence that
of Q(L)
on L is known
(I), and so we deduce
from (65) with rl=
1
(68a)
where h(.r, J*) = - 4~ e Ii ‘1. We may rewrite
From
this integral
formula
this as
for the surface tension, we can reasonably argue that only for 1-v~ ~‘1< 1, but for e-i’- ‘1 is appreciable
0”’ > 0. The exponential 1.~-.I/ < I the entire integrand
in (68b) is positive and hence we expect gXxiito be to positive. Indeed in the strict limit ;‘+O. CT”’diverges to infinity corresponding the infinitely long-ranged attractive potentials”), but this divergence does not concern us here. ii) T/w planur inte[f:fuce in u tlt‘o -c’on~ponent penerruhk -.vplwc tmm’~~i We consider a two-component system with species 1 and 2, in which particles of the same species do not interact while particles of different species interact as hard spheres of radius 1. This is the primitive version of the penetrable sphere model introduced by Widom and Rowlinson”). For a given range of chemical potential ,L /L = ,u’ = ,u’, coexistence of an r-phase, rich in species 1. with a P-phase. rich in species 2, is possible. The structure and thermodynamics of a planar interface between these phases has been extensively studied by Rowlinson et al.“). For a two-phase system with a planar interface (fig. 2) the grand potential
THERMODYNAMIC
PROPERTIES
OF INHOMOGENEOUS
per unit area Q(L) in a mean-field approximation
--=
1kT
has the functional
421
form2)
L+I
L
w,)vo
FLUIDS
dx
s
-L
+
s
-CL+
dy hi& Y)P’(x)P’(Y) I)
1dx ,W)b(d(x)lJ)
- p’(x)1,
(69)
with h,, = hz2= 0 and
h&3y)= hA%Y)= i
lx-Yl>L
0, $1
_
(x
_y)‘],
(x
_
y(
<
1
.
(70)
J. is the common value of the activity for species 1 and 2. The activity 2 and the densities pi and p2 are all dimensionless, having been scaled by the volume r0 = 47r13/3;the length L and coordinate x are also dimensionless, having been scaled by the hard-core diameter 1. Applying the method of variation under extension to the functional (69) just as it was applied (67) in the van der Waals fluid, we find
x+=s s d%l
dx
dy h,(x> y)[p’(x)p’(y)
+
P~(x)PY.Y)I 7
(71)
where
0, h,(x,y)=
i $3(x -1’)l-
lx-,++ 11, IX -yJ < 1 .
(72)
The formula (71) for a@ is identical (under the approximate transcription of the two-component primitive model to the one-component penetrable-sphere model) to the surface-tension formula derived, in a mean-field approximation, by Leng et a1.z3)directly through the canonical partition function. This is not surprising since the method of variation under extension is not unlike the method employed by Leng et a1.23) of scaling the linear dimensions of the system in order to differentiate the canonical partition function with respect to the interfacial area. On the other hand, the agreement between the surface-tension formula of Leng et a1.23)and (71) may be due to the high degree of internal consistency of the present mean-field approximation for the mode122). It is not obvious that the surface tension derived by a differentiation of the canonical free energy with respect to the interfacial area in a mean-field approximation will be identical to the tension derived by the method of variation under extension applied to a mean-field density functional.
M. BOITEUX
422
iii) The three-phase
line qf‘rontact
In a three-component coexisting
phases
respectively, circular
(c(, /I, y), each
section
in a three-component penetrable-sphere
penetrable-sphere
meet along
cross
AND J. KERINS
of which
a straight
with radius
model,
it is possible
is rich
in the species
line of contactj)
line (figs. 3b, 4) then the excess grand
contact
line, a(L)
= Q(L) +pxL’.
sample
potential
per unit length
has the thermodynamic
line
of the
form (cf. eq. 31) (73)
where chi is the interfacial r is the
is a
of the three-phase
Q(L)+J”‘Lt_z.
and
three
i, i = 1,2. 3.
as in fig. (2d). If S(L)
f. of a cylindrical
contact
model
to have
tension
tension of the
of the planar contact
tij. interface
line.
The
(tij, = L$, 87 or r*)~)
straight
contact-line
in-
homogeneity is two-dimensional (ri = 2). but by considering the excess potential o(L) where the pressure contribution is already removed we have fixed W? = 0. The mean-field density functional for this contact line system is
dy h,, (13 y )u’(x )u
‘(y ) +
XI.1 Y S(I.1
d-d(p). s S(l.1
(74)
The equilibrium densities p’(x) (i = I, 2, 3) depend on the position x in S(L), and along with the common value of the activity 2, have been scaled by the volume z’(,= 47#/3; the length L and coordinates x have been scaled by 1. The interaction kernels h,,(x. y) are zero if i =,j. and if i #,j they depend only on z = IX -yl in the form 0, q&Y)
= k,(=)
The function
g(p)
g(p)
=
3 &I
:>I, (75)
of the densities
= P’(x)[ln(p’(x)/i)
-I < 1
-?)“?,
is
- p’(.u)] + $T.
(76)
where p is the common value of the pressure in each of the bulk applying the method of variation under extension we find
xa”“L
=
dx s s S(L) X SC/.+ I)
i i ,:lk=l i#h
P’(x)P”(Y)
1s +
phases.
dx 2$(p),
SC L)
On
(77a)
THERMODYNAMIC
where, with z = IX -
PROPERTIES
OF INHOMOGENEOUS
FLUIDS
423
y(,
(77b)
and
h,(x, y) = h”(Z) =
3 &2;:;;,,2]l
z < ‘.
Note that h,(i) and h,(i), although integrable, are divergent at z = 1; this is a reflection of the underlying hard-sphere interaction. Thus in the functional (74) for the excess grand potential we have distinguished the surface tension contribution
(77) from the line tension
contribution
(78).
iv) Fluid ugainst a solid wall We now generalize our earlier definition of an inhomogeneous the case of a single-component fluid next to a solid. We consider
fluid to include the case in which
the solid-fluid interface is planar, with the fluid density p depending only on the coordinate .Y measured perpendicular to the interface (fig. 7). Furthermore, we model the solid as a rigid, impermeable continuum of density p.!, which fills the entire half-space x < 0. We isolate a macroscopic sample of this solid-fluid system in an imaginary rectangular prism such that the solid-fluid interface is parallel to
i
Fig. 7. The planar solid-fluid interface
M. BOITEUX
424
a pair of opposite 21,. The grand naturally
AND J. KERINS
prism faces of area A and also bisects four prism edges of length potential
per unit
interfacial
area
Q,(L)
of the sample
splits
into two pieces (79)
Q,(L) = Q,,,,,,(L) + Q(L) Q,,,,,,(L) is the contribution
to Q,(L) from the solid in the half of the sample lying
in the region .v < 0, and Q(L) is the contribution the sample in the region .Y > 0. For our purposes, supposed to include any contributions induced by the model solid; Q(L)=
-pL
to Q,(f,)
from the fluid in that half of only Q(L) is of interest and is from
fluid
inhomogeneities
+o.
(80)
where p is the pressure in the bulk fluid and 0 is the solid-fluid interfacial tension. We assume that the interactions of the solid with the fluid are represented by an external field U,(s) acting on the fluid in the region I > 0. The equilibrium density functional for the sample can be written as i
/ C’(L) =
ds I 0
if we assume
/ d.r /I(\-, .r)p(.~)p(~)
s 0
+
d.x G(p) + j_ 0
that the fluid-fluid
interactions
ds U,(s)~>(s)
.
(81)
s /I are well-represented
by the first two
terms on the r.h.s. of (81). For example, in a van der Waals fluid Ii/(p) is the density of grand potential in a hard-sphere reference system and 11(x. .I’) is related to the long-ranged attractive potential’“). Q is now a functional of the density /I(.\-). the interaction kernel Iz(.Y,,I.) and the external field due to the solid U&s). Accordingly, in the variation of Q(L) under extension in L, there will be a term arising from the variation of the external field after the appropriate coordinate transformation.
+
s ds--
6Q
SU,(s)
(SU,(s) + higher-order
where, as always, 652/6p(s) known (63. 64), and
terms in the variations,
= 0. The term in (82) from the variation
(82)
in h(.~, JS) is
(83)
THERMODYNAMIC
PROPERTIES
On the basis of the thermodynamic method
of variation
under
OF INHOMOGENEOUS
equation
extension,
(SO), we can isolate,
the solid-fluid
interfacial
0
0
.i0
through
tension
h(x,y)+;+yE + dxp(x) 8.~1
425
FLUIDS
-xdU’ [
the
0,
.
dx]
(84) From (84) there are clearly distinct contributions to (T from the fluid-fluid interactions and the solid-fluid interactions. Suppose we choose a van der Waals fluid (6668) near a van der Waals solid19) for which U,(x) = -c,
ee‘ .
(85)
In that case we have (cf. eq. 68) 3c
d_v (1 -(x
dx
ya =;
s II
X
3( -Yl)e-I”-~lp(y)p(x)-~,
s
dxx
e-‘p(x).
(86)
s
0
0
Thus, the solid-fluid interactions make a negative contribution to the tension (T, as expected because of the attractive nature of U,. And depending on the magnitude of c, relative to Q, we might easily calculate negative solid-fluid tensions24). We must remark that a negative solid-fluid tension is possible only because of the rigidity of the solid. Consider an interface of tension 0 and with area A in a two-phase equilibrium system in which each bulk phase has a fixed volume. A variation to a nearby equilibrium state, which increases the interfacial area to A + 6.4, must be accompanied by a rearrangement of the material in both of the bulk phases adjacent to the interface, in order to satisfy the constraints of fixed bulk-phase
volume.
Accordingly,
for the supposed
variation,
the total
grand
potential of the system will increase by an amount aL5A due to the interfacial variation and by an amount 6 W due to the necessary rearrangement of the bulk phase materials. Now the material in a fluid phase is so mobile and malleable that no work is done in rearranging the fluid and 6 W = 0 for a bulk fluid phase. In a system of two fluid phases, a variation 6A in the interfacial area causes a variation oaA in grand potential of the system; thus a necessary condition for a stable fluid interface is cr > 0. On the other hand, the material in a bulk solid phase is so rigid that a rearrangement of the solid requires stress-strain work with 6 W > 0. At a solid-fluid interface a variation of 6A in the interfacial area causes a variation o6A + 6 W in the grand potential. The necessary condition of stability for the solid-fluid interface is that adA + 6 W be positive for variations from our
M. BOITEUX
426
initial equilibrium solidNluid
state. Clearly
tensions
AND J. KERINS
this does not imply 0 > 0 in general,
are thermodynamically
and negative
allowed.
5. Discussion In the density variation dynamic
functional
formalism
of inhomogeneous
fluids,
the method
of
under extension is useful for expressing the macroscopic, thermoproperties of the inhomogeneities as functionals of the microscopic, local
densities. In particular, the microscopic expression for the excess grand potential associated with an inhomogeneity follows from a comparison of the variation of the (reduced) grand potential Q(L) under extension in L as computed from a thermodynamic equation (17) with the variation computed from a densityfunctional equation (65). The key step in computing the variation of the density-functional expression for Q(L) is a transformation of coordinates after which the extended system is defined over the same domain of integration as the original unextended system. Although we have restricted our attention to the excess grand potential, the other thermodynamic properties of the inhomogeneity can be generated by differentiation of the excess grand potential with respect to the thermodynamic parameters T and p’, i = 1,2, . . , C (cf. eq. 16, the Gibbs adsorption equation for a planar interface). The method of variation under extension is very general and may be applied to functionals other than Sz. For example, the method could be applied to the functional N’[p’, L] for the (reduced) number of particles of species i in S(L) N’[ p’; L] =
dx p ‘(x )
(87)
to get adsorptions directly. method as presently applied space (i.e. the thermodynamic
The single physical assumption necessary for the is that, although the inhomogeneous fluid fills all limit has already been taken), we can focus our
SC/. ,
attention on a large but finite-sized sample, in which the inhomogeneities are located. Therefore in the thermodynamic equation (17) for L?(L) we allow for terms of order (l/L) due to the finite size of the sample, but we need not allow for other terms which would be associated with edge effects if the fluid itself filled only a finite region of space. The method of variation under extension may also be applied to a much broader class of density functionals than those considered above in sections 3 and 4. For example, the method could be applied to density functionals of the type devised by Nordholm et al.25,‘6) on a phenomenological basis, or to functionals of the type proposed by Percus”) on the basis of exact results for one-dimensional fluids. In
THERMODYNAMIC
both cases the grand p(x),
potential
but also on a function
The method Ebner,
of the density
As a final point,
OF INHOMOGENEOUS
is functionally
in that case the interaction
421
FLUIDS
not only on the density of the density
to the approximate-density-functional
p(x).
model
of
kernel h(x, y) is itself a
p(n).
we note that our approach
extension of the density analysis of deformation
dependent
p(x) which itself is a functional
is also applicable
Saam and Stroud2’);
functional
PROPERTIES
for evaluating
the variation
under
functional Q[p’; L] has a classical antecedent in the in continuous media29). Let us associate the original
sample of inhomogeneous fluid with a continuum of material in an initial or undeformed state which occupies a region S(L) of space. After the appropriate resealing
coordinate
transformation
(24) we may associate
the extended
sample
with the same continuum of material occupying the region S(L) but now in a final or deformed state’“). A pointy in the original sample is shifted to the new point x = x(y, t) associated
under with
deformed states. density functional transport
the
deformation.
a distortion
A variation
of the volume
in the
element
metric
between
tensor
g,, is
the initial
and
Moreover, our calculation of 652[p’; L], the variation of the under extension, may be regarded as an application of Reynolds
theorem”),
6Q[p’, L] = 6 j+dU’(y)= Xl.)
1 dy W’ + ‘f’(V.~(y))l,
(88)
.-WI
with
for a variation feature
with respect to 6 of the local density of grand potential
central
macroscopic, macroscopic microscopic
to the method thermodynamic and microscopic
expressions
of variation equation equation
under
extension
(17) for Q. Only because for O(L)
for the thermodynamic
Y. The novel
is the additional we have both a
are we able to derive
properties
simple
of the inhomogeneities.
The application of the method of variation under extension to an exact density functional which involves the direct correlation function’.‘4.‘2), and the application of the method directly to the grand partition function are currently under investigation”). In these exact microscopic theories for the grand potential Q, the stress-strain interpretation of the variation 652 can be more precisely discussed in terms of the microscopic stress tenso@). The mechanical interpretation of 652 as stress-strain work has its limitations, as recently pointed out in the case of a spherical interface34,35). But whether the microscopic expression for Sz comes from an exact statistical-mechanical theory or an approximate density functional theory, the variation 6Q can be calculated unambiguously on the basis of the
428
M. BOITEUX
macroscopic,
thermodynamic
dent of any detailed
equation
microscopic
is therefore
independent
microscopic
interpretation.
of any
AND
J. KERINS
for Q. Hence the variation
interpretation,
652 is indepen-
such as stress-strain
of the difficulties
associated
with
work,
and
a detailed
Acknowledgements We wish to acknowledge support from the Centre National de la Recherche Scientifique (M.B.) and the U.S. Department of Energy (J.K.). We are grateful to Professors H.T. Davis, M.E. Fisher, L.E. Striven and B. Widom for critical comments. We also appreciate several useful remarks on the spherical interface by A.H. Falls.
References I) 2) 3) 4) 5) 6) 7) 8) 9) 10) 11) 12) 13) 14) 15) 16) 17) 18) 19) 20) 21) 22) 23) 24) 25) 26) 27) 28)
J.W. Gibbs, Collected Works (Longmans and Green, New York, 1928) Vol. I, pp. 55-353. J. Kerins and B. Widom, J. Chem. Phys. 77 (1982) 2061. R. Evans, Adv. Phys. 28 (1979) 143. J. Kerins and M. Boiteux, Physica 117A (1983) 575. B. Widom in: Statistical Mechanics and Statistical Methods in Theory and Application, U. Landmann, ed. (Plenum, New York, 1977) pp. 33371. H.T. Davis and L.E. Striven, Adv. Chem. Phys. 49 (1982) 357. L. Boruvka and A.W. Neumann, J. Chem. Phys. 66 (1977) 5464. IS. Sokolnikoff, Tensor Analysis Theory and Applications (Wiley, New York. 1951). D.E. Soper, Classical Field Theory (Wiley, New York, 1976). L.D. Landau and E.M. Lifshitz, The Classical Theory of Fields (Pergamon, Oxford, 1975) fourth revised English edition. Ref. IO, p. 273. J. Kerins, Ph.D. Thesis, Cornell University, 1982. J.S. Rowlinson and B. Widom, Molecular Theory of Capillarity (University Press, Oxford. 1982). A.J.M. Yang, P.D. Fleming and J.H. Gibbs, J. Chem. Phys. 64 (1976) 3732. A.H. Falls, L.E. Striven and H.T. Davis, J. Chem. Phys. 75 (1981) 3986. N.G. van Kampen, Phys. Rev. A 135 (1964) 362. C. Varea, A. Valderama and A. Robledo, J. Chem. Phys. 73 (1980) 6265. A. Robiedo and C. Varea, J. Stat. Phys. 26 (1981) 513. D.E. Sullivan, J. Chem. Phys. 74 (1981) 2604. M. Kac, Phys. Fluids 2 (1959) 8; M. Kac, G. Uhlenbeck and P.C. Hemmer, J. Math. Phys. 4 (1963) 216. B. Widom and J.S. Rowlinson, J. Chem. Phys. 52 (1970) 1670. J.S. Rowlinson, Adv. Chem. Phys. 41 (1980) 2 and references therein. CA. Leng, J.S. Rowlinson and S.M. Thompson, Proc. Roy. Sot. A352 (1976) I; A358 (1977) 267. G.F. Teletzke, Ph.D. Thesis, University of Minnesota, 1983. S. Nordholm, M. Johnson and B.C. Freasier, Aust. J. Chem. (to be published). M. Johnson and S. Nordholm, J. Chem. Phys. 75 (1981) 1953. J.K. Percus, J. Chem. Phys. 75 (1981) 1316. C. Ebner, W.F. Saam and D. Stroud, Phys. Rev. A 14 (1976) 2264.
THERMODYNAMIC
29) We are grateful 30) 3 I) 32) 33) 34) 35)
to Prof.
PROPERTIES
L.E. Striven
OF INHOMOGENEOUS
for pointing
out the connection
FLUIDS
between
the method
429
of
variation under extension and Reynolds transport theorem. Ref. 8, Ch. 6. R. Ark, Vectors, Tensors and the Basic Equations of Fluid Mechanics (Prentice-Hall, Englewood Cliffs, 1962) p. 84. W.F. Saam and C. Ebner, Phys. Rev. A 15 (1977) 2566. J. Kerins (unpublished). P. Schofield and J.R. Henderson, Proc. R. Sot. Lond. A 379 (1982) 231. S.J. Hemingway, J.R. Henderson and J.S. Rowlinson in: Structure of the Interfacial Region, Faraday
Symp.
No. I6 (Royal
Society
of Chemistry,
London,
1982)
pp. l-l I.